首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  相似文献   

4.
5.
Lysophosphatidic acid (LPA), a bioactive phospholipid, induces a wide range of cellular effects, including gene expression, cytoskeletal rearrangement, and cell survival. We have previously shown that LPA stimulates secretion of pro- and anti-inflammatory cytokines in bronchial epithelial cells. This study provides evidence that LPA enhances pulmonary epithelial barrier integrity through protein kinase C (PKC) δ- and ζ-mediated E-cadherin accumulation at cell-cell junctions. Treatment of human bronchial epithelial cells (HBEpCs) with LPA increased transepithelial electrical resistance (TER) by ∼2.0-fold and enhanced accumulation of E-cadherin to the cell-cell junctions through Gαi-coupled LPA receptors. Knockdown of E-cadherin with E-cadherin small interfering RNA or pretreatment with EGTA (0.1 mm) prior to LPA (1 μm) treatment attenuated LPA-induced increases in TER in HBEpCs. Furthermore, LPA induced tyrosine phosphorylation of focal adhesion kinase (FAK) and overexpression of the FAK inhibitor, and FAK-related non-kinase-attenuated LPA induced increases in TER and E-cadherin accumulation at cell-cell junctions. Overexpression of dominant negative protein kinase δ and ζ attenuated LPA-induced phosphorylation of FAK, accumulation of E-cadherin at cell-cell junctions, and an increase in TER. Additionally, lipopolysaccharide decreased TER and induced E-cadherin relocalization from cell-cell junctions to cytoplasm in a dose-dependent fashion, which was restored by LPA post-treatment in HBEpCs. Intratracheal post-treatment with LPA (5 μm) reduced LPS-induced neutrophil influx, protein leak, and E-cadherin shedding in bronchoalveolar lavage fluids in a murine model of acute lung injury. These data suggest a protective role of LPA in airway inflammation and remodeling.The airway epithelium is the site of first contact for inhaled environmental stimuli, functions as a physical barrier to environmental insult, and is an essential part of innate immunity. Epithelial barrier disruption is caused by inhaled allergens, dust, and irritants, resulting in inflammation, bronchoconstriction, and edema as seen in asthma and other respiratory diseases (14). Furthermore, increased epithelial permeability also results in para-cellular leakage of large proteins, such as albumin, immunoglobulin G, and polymeric immunoglobulin A, into the airway lumen (5, 6). The epithelial cell-cell junctional complex is composed of tight junctions, adherens junctions, and desmosomes. These adherens junctions play a pivotal role in regulating the activity of the entire junctional complex because the formation of adherens junctions subsequently leads to the formation of other cell-cell junctions (79). The major adhesion molecules in the adherens junctions are the cadherins. E-cadherin is a member of the cadherin family that mediates calcium-dependent cell-cell adhesion. The N-terminal ectodomain of E-cadherin contains homophilic interaction specificity, and the cytoplasmic domain binds to catenins, which interact with actin (1013). Plasma membrane localization of E-cadherin is critical for the maintenance of epithelial cell-cell junctions and airway epithelium integrity (7, 10, 14). A decrease of adhesive properties of E-cadherin is related to the loss of differentiation and the subsequent acquisition of a higher motility and invasiveness of epithelial cells (10, 14, 15). Dislocation or shedding of E-cadherin in the airway epithelium induces epithelial shedding and increases airway permeability in lung airway diseases (10, 14, 16). In an ovalbumin-challenged guinea pig model of asthma, it has been demonstrated that E-cadherin is dislocated from the lateral margins of epithelial cells (10). Histamine increases airway para-cellular permeability and results in an increased susceptibility of airway epithelial cells to adenovirus infection by interrupting E-cadherin adhesion (14). Serine phosphorylation of E-cadherin by casein kinase II, GSK-3β, and PKD1/PKC2 μ enhanced E-cadherin-mediated cell-cell adhesion in NIH3T3 fibroblasts and LNCaP prostate cancer cells (11, 17). However, the regulation and mechanism by which E-cadherin is localized within the pulmonary epithelium is not fully known, particularly during airway remodeling.LPA, a naturally occurring bioactive lipid, is present in body fluids, such as plasma, saliva, follicular fluid, malignant effusions, and bronchoalveolar lavage (BAL) fluids (1820). Six distinct high affinity cell-surface LPA receptors, LPA-R1–6, have been cloned and described in mammals (2126). Extracellular activities of LPA include cell proliferation, motility, and cell survival (2730). LPA exhibits a wide range of effects on differing cell types, including pulmonary epithelial, smooth muscle, fibroblasts, and T cells (3135). LPA augments migration and cytokine synthesis in lymphocytes and induces chemotaxis of Jurkat T cells through Matrigel membranes (34). LPA induces airway smooth muscle cell contractility, proliferation, and airway repair and remodeling (35, 36). LPA also potently stimulates IL-8 (31, 3739), IL-13 receptor α2 (IL-13Rα2) (40), and COX-2 gene expression and prostaglandin E2 release (41) in HBEpCs. Prostaglandin E2 and IL-13Rα2 have anti-inflammatory properties in pulmonary inflammation (42, 43). These results suggest that LPA may play a protective role in lung disease by stimulating an innate immune response while simultaneously attenuating the adaptive immune response. Furthermore, intravenous injection with LPA attenuated bacterial endotoxin-induced plasma tumor necrosis factor-α production and myeloperoxidase activity in the lungs of mice (44), suggesting an anti-inflammatory role for LPA in a murine model of sepsis.We have reported that LPA induces E-cadherin/c-Met accumulation in cell-cell contacts and increases TER in HBEpCs (45). Here, for the first time, we report that LPA-induced increases in TER are dependent on PKCδ, PKCζ, and FAK-mediated E-cadherin accumulation at cell-cell junctions. Furthermore, we demonstrate that post-treatment of LPA rescues LPS-induced airway epithelial disruption in vitro and reduces E-cadherin shedding in a murine model of ALI. This study identifies the molecular mechanisms linking the LPA and LPA receptors to maintaining normal pulmonary epithelium barrier function, which is critical in developing novel therapies directed at ameliorating pulmonary diseases.  相似文献   

6.
Prion propagation involves a conformational transition of the cellular form of prion protein (PrPC) to a disease-specific isomer (PrPSc), shifting from a predominantly α-helical conformation to one dominated by β-sheet structure. This conformational transition is of critical importance in understanding the molecular basis for prion disease. Here, we elucidate the conformational properties of a disulfide-reduced fragment of human PrP spanning residues 91–231 under acidic conditions, using a combination of heteronuclear NMR, analytical ultracentrifugation, and circular dichroism. We find that this form of the protein, which similarly to PrPSc, is a potent inhibitor of the 26 S proteasome, assembles into soluble oligomers that have significant β-sheet content. The monomeric precursor to these oligomers exhibits many of the characteristics of a molten globule intermediate with some helical character in regions that form helices I and III in the PrPC conformation, whereas helix II exhibits little evidence for adopting a helical conformation, suggesting that this region is a likely source of interaction within the initial phases of the transformation to a β-rich conformation. This precursor state is almost as compact as the folded PrPC structure and, as it assembles, only residues 126–227 are immobilized within the oligomeric structure, leaving the remainder in a mobile, random-coil state.Prion diseases, such as Creutzfeldt-Jacob and Gerstmann-Sträussler-Scheinker in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle, are fatal neurological disorders associated with the deposition of an abnormally folded form of a host-encoded glycoprotein, prion (PrP)2 (1). These diseases may be inherited, arise sporadically, or be acquired through the transmission of an infectious agent (2, 3). The disease-associated form of the protein, termed the scrapie form or PrPSc, differs from the normal cellular form (PrPC) through a conformational change, resulting in a significant increase in the β-sheet content and protease resistance of the protein (3, 4). PrPC, in contrast, consists of a predominantly α-helical structured domain and an unstructured N-terminal domain, which is capable of binding a number of divalent metals (512). A single disulfide bond links two of the main α-helices and forms an integral part of the core of the structured domain (13, 14).According to the protein-only hypothesis (15), the infectious agent is composed of a conformational isomer of PrP (16) that is able to convert other isoforms to the infectious isomer in an autocatalytic manner. Despite numerous studies, little is known about the mechanism of conversion of PrPC to PrPSc. The most coherent and general model proposed thus far is that PrPC fluctuates between the dominant native state and minor conformations, one or a set of which can self-associate in an ordered manner to produce a stable supramolecular structure composed of misfolded PrP monomers (3, 17). This stable, oligomeric species can then bind to, and stabilize, rare non-native monomer conformations that are structurally complementary. In this manner, new monomeric chains are recruited and the system can propagate.In view of the above model, considerable effort has been devoted to generating and characterizing alternative, possibly PrPSc-like, conformations in the hope of identifying common properties or features that facilitate the formation of amyloid oligomers. This has been accomplished either through PrPSc-dependent conversion reactions (1820) or through conversion of PrPC in the absence of a PrPSc template (2125). The latter approach, using mainly disulfide-oxidized recombinant PrP, has generated a wide range of novel conformations formed under non-physiological conditions where the native state is relatively destabilized. These conformations have ranged from near-native (14, 26, 27), to those that display significant β-sheet content (21, 23, 2833). The majority of these latter species have shown a high propensity for aggregation, although not all are on-pathway to the formation of amyloid. Many of these non-native states also display some of the characteristics of PrPSc, such as increased β-sheet content, protease resistance, and a propensity for oligomerization (28, 29, 31) and some have been claimed to be associated with the disease process (34).One such PrP folding intermediate, termed β-PrP, differs from the majority of studied PrP intermediate states in that it is formed by refolding the PrP molecule from the native α-helical conformation (here termed α-PrP), at acidic pH in a reduced state, with the disulfide bond broken (22, 35). Although no covalent differences between the PrPC and PrPSc have been consistently identified to date, the role of the disulfide bond in prion propagation remains disputed (25, 3639). β-PrP is rich in β-sheet structure (22, 35), and displays many of the characteristics of a PrPSc-like precursor molecule, such as partial resistance to proteinase K digestion, and the ability to form amyloid fibrils in the presence of physiological concentrations of salts (40).The β-PrP species previously characterized, spanning residues 91–231 of PrP, was soluble at low ionic strength buffers and monomeric, according to elution volume on gel filtration (22). NMR analysis showed that it displayed radically different spectra to those of α-PrP, with considerably fewer observable peaks and markedly reduced chemical shift dispersion. Data from circular dichroism experiments showed that fixed side chain (tertiary) interactions were lost, in contrast to the well defined β-sheet secondary structure, and thus in conjunction with the NMR data, indicated that β-PrP possessed a number of characteristics associated with a “molten globule” folding intermediate (22). Such states have been proposed to be important in amyloid and fibril formation (41). Indeed, antibodies raised against β-PrP (e.g. ICSM33) are capable of recognizing native PrPSc (but not PrPC) (4244). Subsequently, a related study examining the role of the disulfide bond in PrP folding confirmed that a monomeric molten globule-like form of PrP was formed on refolding the disulfide-reduced protein at acidic pH, but reported that, under their conditions, the circular dichroism response interpreted as β-sheet structure was associated with protein oligomerization (45). Indeed, atomic force microscopy on oligomeric full-length β-PrP (residues 23–231) shows small, round particles, showing that it is capable of formation of oligomers without forming fibrils (35). Notably, however, salt-induced oligomeric β-PrP has been shown to be a potent inhibitor of the 26 S proteasome, in a similar manner to PrPSc (46). Impairment of the ubiquitin-proteasome system in vivo has been linked to prion neuropathology in prion-infected mice (46).Although the global properties of several PrP intermediate states have been determined (3032, 35), no information on their conformational properties on a sequence-specific basis has been obtained. Their conformational properties are considered important, as the elucidation of the chain conformation may provide information on the way in which these chains pack in the assembly process, and also potentially provide clues on the mechanism of amyloid assembly and the phenomenon of prion strains. As the conformational fluctuations and heterogeneity of molten globule states give rise to broad NMR spectra that preclude direct observation of their conformational properties by NMR (4750), here we use denaturant titration experiments to determine the conformational properties of β-PrP, through the population of the unfolded state that is visible by NMR. In addition, we use circular dichroism and analytical ultracentrifugation to examine the global structural properties, and the distribution of multimeric species that are formed from β-PrP.  相似文献   

7.
8.
9.
10.
11.
The present study tests the hypothesis that the structure of extracellular domain Loop 2 can markedly affect ethanol sensitivity in glycine receptors (GlyRs) and γ-aminobutyric acid type A receptors (GABAARs). To test this, we mutated Loop 2 in the α1 subunit of GlyRs and in the γ subunit of α1β2γ2GABAARs and measured the sensitivity of wild type and mutant receptors expressed in Xenopus oocytes to agonist, ethanol, and other agents using two-electrode voltage clamp. Replacing Loop 2 of α1GlyR subunits with Loop 2 from the δGABAAR (δL2), but not the γGABAAR subunit, reduced ethanol threshold and increased the degree of ethanol potentiation without altering general receptor function. Similarly, replacing Loop 2 of the γ subunit of GABAARs with δL2 shifted the ethanol threshold from 50 mm in WT to 1 mm in the GABAA γ-δL2 mutant. These findings indicate that the structure of Loop 2 can profoundly affect ethanol sensitivity in GlyRs and GABAARs. The δL2 mutations did not affect GlyR or GABAAR sensitivity, respectively, to Zn2+ or diazepam, which suggests that these δL2-induced changes in ethanol sensitivity do not extend to all allosteric modulators and may be specific for ethanol or ethanol-like agents. To explore molecular mechanisms underlying these results, we threaded the WT and δL2 GlyR sequences onto the x-ray structure of the bacterial Gloeobacter violaceus pentameric ligand-gated ion channel homologue (GLIC). In addition to being the first GlyR model threaded on GLIC, the juxtaposition of the two structures led to a possible mechanistic explanation for the effects of ethanol on GlyR-based on changes in Loop 2 structure.Alcohol abuse and dependence are significant problems in our society, with ∼14 million people in the United States being affected (1, 2). Alcohol causes over 100,000 deaths in the United States, and alcohol-related issues are estimated to cost nearly 200 billion dollars annually (2). To address this, considerable attention has focused on the development of medications to prevent and treat alcohol-related problems (35). The development of such medications would be aided by a clear understanding of the molecular structures on which ethanol acts and how these structures influence receptor sensitivity to ethanol.Ligand-gated ion channels (LGICs)2 have received substantial attention as putative sites of ethanol action that cause its behavioral effects (612). Research in this area has focused on investigating the effects of ethanol on two large superfamilies of LGICs: 1) the Cys-loop superfamily of LGICs (13, 14), whose members include nicotinic acetylcholine, 5-hydroxytryptamine3, γ-aminobutyric acid type A (GABAA), γ-aminobutyric acid type C, and glycine receptors (GlyRs) (10, 11, 1520) and 2) the glutamate superfamily, including N-methyl d-aspartate, α-amino-3-hydroxyisoxazolepropionic acid, and kainate receptors (21, 22). Recent studies have also begun investigating ethanol action in the ATP-gated P2X superfamily of LGICs (2325).A series of studies that employed chimeric and mutagenic strategies combined with sulfhydryl-specific labeling identified key regions within Cys-loop receptors that appear to be initial targets for ethanol action that also can determine the sensitivity of the receptors to ethanol (712, 18, 19, 2630). This work provides several lines of evidence that position 267 and possibly other sites in the transmembrane (TM) domain of GlyRs and homologous sites in GABAARs are targets for ethanol action and that mutations at these sites can influence ethanol sensitivity (8, 9, 26, 31).Growing evidence from GlyRs indicates that ethanol also acts on the extracellular domain. The initial findings came from studies demonstrating that α1GlyRs are more sensitive to ethanol than are α2GlyRs despite the high (∼78%) sequence homology between α1GlyRs and α2GlyRs (32). Further work found that an alanine to serine exchange at position 52 (A52S) in Loop 2 can eliminate the difference in ethanol sensitivity between α1GlyRs and α2GlyRs (18, 20, 33). These studies also demonstrated that mutations at position 52 in α1GlyRS and the homologous position 59 in α2GlyRs controlled the sensitivity of these receptors to a novel mechanistic ethanol antagonist (20). Collectively, these studies suggest that there are multiple sites of ethanol action in α1GlyRs, with one site located in the TM domain (e.g. position 267) and another in the extracellular domain (e.g. position 52).Subsequent studies revealed that the polarity of the residue at position 52 plays a key role in determining the sensitivity of GlyRs to ethanol (20). The findings with polarity in the extracellular domain contrast with the findings at position 267 in the TM domain, where molecular volume, but not polarity, significantly affected ethanol sensitivity (9). Taken together, these findings indicate that the physical-chemical parameters of residues at positions in the extracellular and TM domains that modulate ethanol effects and/or initiate ethanol action in GlyRs are not uniform. Thus, knowledge regarding the physical-chemical properties that control agonist and ethanol sensitivity is key for understanding the relationship between the structure and the actions of ethanol in LGICs (19, 31, 3440).GlyRs and GABAARs, which differ significantly in their sensitivities to ethanol, offer a potential method for identifying the structures that control ethanol sensitivity. For example, α1GlyRs do not reliably respond to ethanol concentrations less than 10 mm (32, 33, 41). Similarly, γ subunit-containing GABAARs (e.g. α1β2γ2), the most predominantly expressed GABAARs in the central nervous system, are insensitive to ethanol concentrations less than 50 mm (42, 43). In contrast, δ subunit-containing GABAARs (e.g. α4β3δ) have been shown to be sensitive to ethanol concentrations as low as 1–3 mm (4451). Sequence alignment of α1GlyR, γGABAAR, and δGABAAR revealed differences between the Loop 2 regions of these receptor subunits. Since prior studies found that mutations of Loop 2 residues can affect ethanol sensitivity (19, 20, 39), the non-conserved residues in Loop 2 of GlyR and GABAAR subunits could provide the physical-chemical and structural bases underlying the differences in ethanol sensitivity between these receptors.The present study tested the hypothesis that the structure of Loop 2 can markedly affect the ethanol sensitivity of GlyRs and GABAARs. To accomplish this, we performed multiple mutations that replaced the Loop 2 region of the α1 subunit in α1GlyRs and the Loop 2 region of the γ subunit of α1β2γ2 GABAARs with corresponding non-conserved residues from the δ subunit of GABAAR and tested the sensitivity of these receptors to ethanol. As predicted, replacing Loop 2 of WT α1GlyRs with the homologous residues from the δGABAAR subunit (δL2), but not the γGABAAR subunit (γL2), markedly increased the sensitivity of the receptor to ethanol. Similarly, replacing the non-conserved residues of the γ subunit of α1β2γ2 GABAARs with δL2 also markedly increased ethanol sensitivity of GABAARs. These findings support the hypothesis and suggest that Loop 2 may play a role in controlling ethanol sensitivity across the Cys-loop superfamily of receptors. The findings also provide the basis for suggesting structure-function relationships in a new molecular model of the GlyR based on the bacterial Gloeobacter violaceus pentameric LGIC homologue (GLIC).  相似文献   

12.
AMP-activated protein kinase (AMPK) is a heterotrimeric complex playing a crucial role in maintaining cellular energy homeostasis. Recently, homodimerization of mammalian AMPK and yeast ortholog SNF1 was shown by us and others. In SNF1, it involved specific hydrophobic residues in the kinase domain αG-helix. Mutation of the corresponding AMPK α-subunit residues (Val-219 and Phe-223) to glutamate reduced the tendency of the kinase to form higher order homo-oligomers, as was determined by the following three independent techniques in vitro: (i) small angle x-ray scattering, (ii) surface plasmon resonance spectroscopy, and (iii) two-dimensional blue native/SDS-PAGE. Recombinant protein as well as AMPK in cell lysates of primary cells revealed distinct complexes of various sizes. In particular, the assembly of very high molecular mass complexes was dependent on both the αG-helix-mediated hydrophobic interactions and kinase activation. In vitro and when overexpressed in double knock-out (α1−/−, α2−/−) mouse embryonic fibroblast cells, activation of mutant AMPK was impaired, indicating a critical role of the αG-helix residues for AMPK activation via its upstream kinases. Also inactivation by protein phosphatase 2Cα was affected in mutant AMPK. Importantly, activation of mutant AMPK by LKB1 was restored by exchanging the corresponding and conserved hydrophobic αG-helix residues of LKB1 (Ile-260 and Phe-264) to positively charged amino acids. These mutations functionally rescued LKB1-dependent activation of mutant AMPK in vitro and in cell culture. Our data suggest a physiological role for the hydrophobic αG-helix residues in homo-oligomerization of heterotrimers and cellular interactions, in particular with upstream kinases, indicating an additional level of AMPK regulation.The maintenance of energy homeostasis is a basic requirement of all living organisms. The AMP-activated protein kinase (AMPK)2 is crucially involved in this essential process by playing a central role in sensing and regulating energy metabolism on the cellular and whole body level (16). AMPK is also participating in several signaling pathways associated with cancer and metabolic diseases, like type 2 diabetes mellitus, obesity, and other metabolic disorders (79).Mammalian AMPK belongs to a highly conserved family of serine/threonine protein kinases with homologs found in all eukaryotic organisms examined (1, 3, 10). Its heterotrimeric structure includes a catalytic α-subunit and regulatory β- and γ-subunits. These subunits exist in different isoforms (α1, α2, β1, β2, γ1, γ2, and γ3) and splice variants (for γ2 and γ3) and can thus assemble to a broad variety of heterotrimeric isoform combinations. The α- and β-subunits possess multiple autophosphorylation sites, which have been implicated in regulation of subcellular localization and kinase activation (1115). The most critical step of AMPK activation, however, is phosphorylation of Thr-172 within the activation segment of the α-subunit kinase domain. At least two AMPK upstream kinases (AMPKKs) have been identified so far, namely the tumor suppressor kinase LKB1 in complex with MO25 and STRAD (16) and Ca2+/calmodulin-dependent protein kinase kinase-2 (CamKK2) (17). Furthermore, the transforming growth factor-β-activated kinase 1 was also shown to activate AMPK using a variety of in vitro approaches (18), but the physiological relevance of these findings remains unclear. Besides direct phosphorylation of Thr-172, AMPK activity is stimulated by the allosteric activator AMP, which can bind to two Bateman domains formed by two pairs of CBS domains within the γ-subunit (1922). Hereby bound AMP not only allosterically stimulates AMPK but also protects Thr-172 from dephosphorylation by protein phosphatase 2Cα (PP2Cα) and thus hinders inactivation of the kinase (19, 22, 23). Consequently, on the cellular level, AMPK is activated upon metabolic stress increasing the AMP/ATP ratio. Furthermore, AMPK activation can also be induced by several chemical compounds, like nucleoside 5-aminoimidazole-4-carboxamide-1-β-d-ribofuranoside (24) and the anti-diabetic drug Metformin (2528). In addition, the small molecule compound A-769662 was recently developed as a direct allosteric activator of AMPK (29, 30).Previous work in our groups proposed a model of AMPK regulation by AMP, which incorporates the major functional features and the latest structural information (31). The latter mainly included truncated core complexes of AMPK from different species (3235). Further valuable structural information is provided by the x-ray structures of the isolated catalytic domains, in particular of the human AMPK α2-subunit (Protein Data Bank code 2H6D) and its yeast ortholog SNF1 (36, 37). The kinase domain of SNF1 is capable of forming homodimers in the protein crystal, as well as in vitro in solution, in a unique way, which has not been observed previously in any other kinase (36). The dimer interface is predominantly formed by hydrophobic interactions of the loop-αG region, also known as subdomain X situated on the large kinase lobe (36, 38, 39), and it mainly involves Ile-257 and Phe-261. Because the T-loop activation segment was buried within the dimer interface, it was suggested that the dimeric state of the SNF1 catalytic domain represents the inactive form of the kinase. Intriguingly, it was shown in our groups by small angle x-ray scattering that AMPK self-organizes in a concentration-dependent manner to form homo-oligomers in solution (31). However, the interface responsible for oligomerization of the AMPK heterotrimer has remained elusive.Here we further investigate the distinct oligomeric states of the AMPK heterotrimer and suggest a possible regulatory function for this process. Most importantly, we provide conclusive evidence for participation of αG-helix residues in the recognition of AMPK by its upstream kinases LKB1 and CamKK2.  相似文献   

13.
14.
Understanding the structural and assembly dynamics of the amyloid β-protein (Aβ) has direct relevance to the development of therapeutic agents for Alzheimer disease. To elucidate these dynamics, we combined scanning amino acid substitution with a method for quantitative determination of the Aβ oligomer frequency distribution, photo-induced cross-linking of unmodified proteins (PICUP), to perform “scanning PICUP.” Tyr, a reactive group in PICUP, was substituted at position 1, 10, 20, 30, or 40 (for Aβ40) or 42 (for Aβ42). The effects of these substitutions were probed using circular dichroism spectroscopy, thioflavin T binding, electron microscopy, PICUP, and mass spectrometry. All peptides displayed a random coil → α/β → β transition, but substitution-dependent alterations in assembly kinetics and conformer complexity were observed. Tyr1-substituted homologues of Aβ40 and Aβ42 assembled the slowest and yielded unusual patterns of oligomer bands in gel electrophoresis experiments, suggesting oligomer compaction had occurred. Consistent with this suggestion was the observation of relatively narrow [Tyr1]Aβ40 fibrils. Substitution of Aβ40 at the C terminus decreased the population conformational complexity and substantially extended the highest order of oligomers observed. This latter effect was observed in both Aβ40 and Aβ42 as the Tyr substitution position number increased. The ability of a single substitution (Tyr1) to alter Aβ assembly kinetics and the oligomer frequency distribution suggests that the N terminus is not a benign peptide segment, but rather that Aβ conformational dynamics and assembly are affected significantly by the competition between the N and C termini to form a stable complex with the central hydrophobic cluster.Alzheimer disease (AD)4 is the most common cause of late-life dementia (1) and is estimated to afflict more than 27 million people worldwide (2). An important etiologic hypothesis is that amyloid β-protein (Aβ) oligomers are the proximate neurotoxins in AD. Substantial in vivo and in vitro evidence supports this hypothesis (312). Neurotoxicity studies have shown that Aβ assemblies are potent neurotoxins (5, 1320), and the toxicity of some oligomers can be greater than that of the corresponding fibrils (21). Soluble Aβ oligomers inhibit hippocampal long term potentiation (4, 5, 13, 15, 17, 18, 22) and disrupt cognitive function (23). Compounds that bind and disrupt the formation of oligomers have been shown to block the neurotoxicity of Aβ (24, 25). Importantly, recent studies in higher vertebrates (dogs) have shown that substantial reduction in amyloid deposits in the absence of decreases in oligomer concentration has little effect on recovery of neurological function (26).Recent studies of Aβ oligomers have sought to correlate oligomer size and biological activity. Oligomers in the supernatants of fibril preparations centrifuged at 100,000 × g caused sustained calcium influx in rat hippocampal neurons, leading to calpain activation and dynamin 1 degradation (27). Aβ-derived diffusible ligand-like Aβ42 oligomers induced inflammatory responses in cultured rat astrocytes (28). A 90-kDa Aβ42 oligomer (29) has been shown to activate ERK1/2 in rat hippocampal slices (30) and bind avidly to human cortical neurons (31), in both cases causing apoptotic cell death. A comparison of the time dependence of the toxic effects of the 90-kDa assembly with that of Aβ-derived diffusible ligands revealed a 5-fold difference, Aβ-derived diffusible ligands requiring more time for equivalent effects (31). A 56-kDa oligomer, “Aβ*56,” was reported to cause memory impairment in middle-aged transgenic mice expressing human amyloid precursor protein (32). A nonamer also had adverse effects. Impaired long term potentiation in rat brain slices has been attributed to Aβ trimers identified in media from cultured cells expressing human amyloid precursor protein (33). Dimers and trimers from this medium also have been found to cause progressive loss of synapses in organotypic rat hippocampal slices (10). In mice deficient in neprilysin, an enzyme that has been shown to degrade Aβ in vivo (34), impairment in neuronal plasticity and cognitive function correlated with significant increases in Aβ dimer levels and synapse-associated Aβ oligomers (35).The potent pathologic effects of Aβ oligomers provide a compelling reason for elucidating the mechanism(s) of their formation. This has been a difficult task because of the metastability and polydispersity of Aβ assemblies (36). To obviate these problems, we introduced the use of the method of photo-induced cross-linking of unmodified proteins (PICUP) to rapidly (<1 s) and covalently stabilize oligomer mixtures (for reviews see Refs. 37, 38). Oligomers thus stabilized no longer exist in equilibrium with monomers or each other, allowing determination of oligomer frequency distributions by simple techniques such as SDS-PAGE (37). Recently, to obtain population-average information on contributions to fibril formation of amino acid residues at specific sites in Aβ, we employed a scanning intrinsic fluorescence approach (39). Tyr was used because it is a relatively small fluorophore, exists natively in Aβ, and possesses the side chain most reactive in the PICUP chemistry (40). Using this approach, we found that the central hydrophobic cluster region (Leu17–Ala21) was particularly important in controlling fibril formation of Aβ40, whereas the C terminus was the predominant structural element controlling Aβ42 assembly (39). Here we present results of studies in which key strategic features of the two methods have been combined to enable execution of “scanning PICUP” and the consequent revelation of site-specific effects on Aβ oligomerization.  相似文献   

15.
16.
17.
Paneth cells are a secretory epithelial lineage that release dense core granules rich in host defense peptides and proteins from the base of small intestinal crypts. Enteric α-defensins, termed cryptdins (Crps) in mice, are highly abundant in Paneth cell secretions and inherently resistant to proteolysis. Accordingly, we tested the hypothesis that enteric α-defensins of Paneth cell origin persist in a functional state in the mouse large bowel lumen. To test this idea, putative Crps purified from mouse distal colonic lumen were characterized biochemically and assayed in vitro for bactericidal peptide activities. The peptides comigrated with cryptdin control peptides in acid-urea-PAGE and SDS-PAGE, providing identification as putative Crps. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry experiments showed that the molecular masses of the putative α-defensins matched those of the six most abundant known Crps, as well as N-terminally truncated forms of each, and that the peptides contain six Cys residues, consistent with identities as α-defensins. N-terminal sequencing definitively revealed peptides with N termini corresponding to full-length, (des-Leu)-truncated, and (des-Leu-Arg)-truncated N termini of Crps 1–4 and 6. Crps from mouse large bowel lumen were bactericidal in the low micromolar range. Thus, Paneth cell α-defensins secreted into the small intestinal lumen persist as intact and functional forms throughout the intestinal tract, suggesting that the peptides may mediate enteric innate immunity in the colonic lumen, far from their upstream point of secretion in small intestinal crypts.Antimicrobial peptides (AMPs)2 are released by epithelial cells onto mucosal surfaces as effectors of innate immunity (15). In mammals, most AMPs derive from two major families, the cathelicidins and defensins (6). The defensins comprise the α-, β-, and θ-defensin subfamilies, which are defined by the presence of six cysteine residues paired in characteristic tridisulfide arrays (7). α-Defensins are highly abundant in two primary cell lineages: phagocytic leukocytes, primarily neutrophils, of myeloid origin and Paneth cells, which are secretory epithelial cells located at the base of the crypts of Lieberkühn in the small intestine (810). Neutrophil α-defensins are stored in azurophilic granules and contribute to non-oxidative microbial cell killing in phagolysosomes (11, 12), except in mice whose neutrophils lack defensins (13). In the small bowel, α-defensins and other host defense proteins (1418) are released apically as components of Paneth cell secretory granules in response to cholinergic stimulation and after exposure to bacterial antigens (19). Therefore, the release of Paneth cell products into the crypt lumen is inferred to protect mitotically active crypt cells from colonization by potential pathogens and confer protection against enteric infection (7, 20, 21).Under normal, homeostatic conditions, Paneth cells are not found outside the small bowel, although they may appear ectopically in response to local inflammation throughout the gastrointestinal tract (22, 23). Paneth cell numbers increase progressively throughout the small intestine, occurring at highest numbers in the distal ileum (24). Mouse Paneth cells express numerous α-defensin isoforms, termed cryptdins (Crps) (25), that have broad spectrum antimicrobial activities (6, 26). Collectively, α-defensins constitute approximately seventy percent of the bactericidal peptide activity in mouse Paneth cell secretions (19), selectively killing bacteria by membrane-disruptive mechanisms (2730). The role of Paneth cell α-defensins in gastrointestinal mucosal immunity is evident from studies of mice transgenic for human enteric α-defensin-5, HD-5, which are immune to infection by orally administered Salmonella enterica sv. typhimurium (S. typhimurium) (31).The biosynthesis of mature, bactericidal α-defensins from their inactive precursors requires activation by lineage-specific proteolytic convertases. In mouse Paneth cells, inactive ∼8.4-kDa Crp precursors are processed intracellularly into microbicidal ∼4-kDa Crps by specific cleavage events mediated by matrix metalloproteinase-7 (MMP-7) (32, 33). MMP-7 null mice exhibit increased susceptibility to systemic S. typhimurium infection and decreased clearance of orally administered non-invasive Escherichia coli (19, 32). Although the α-defensin proregions are sensitive to proteolysis, the mature, disulfide-stabilized peptides resist digestion by their converting enzymes in vitro, whether the convertase is MMP-7 (32), trypsin (34), or neutrophil serine proteinases (35). Because α-defensins resist proteolysis in vitro, we hypothesized that Paneth cell α-defensins resist degradation and remain in a functional state in the large bowel, a complex, hostile environment containing varied proteases of both host and microbial origin.Here, we report on the isolation and characterization of a population of enteric α-defensins from the mouse colonic lumen. Full-length and N-terminally truncated Paneth cell α-defensins were identified and are abundant in the distal large bowel lumen.  相似文献   

18.
Rho GTPases are critical components of cellular signal transduction pathways. Both hyperactivity and overexpression of these proteins have been observed in human cancers and have been implicated as important factors in metastasis. We previously showed that dietary n-6 fatty acids increase cancer cell adhesion to extracellular matrix proteins, such as type IV collagen. Here we report that in MDA-MB-435 human melanoma cells, arachidonic acid activates RhoA, and inhibition of RhoA signaling with either C3 exoenzyme or dominant negative Rho blocked arachidonic acid-induced cell adhesion. Inhibition of the Rho kinase (ROCK) with either small molecule inhibitors or ROCK II-specific small interfering RNA (siRNA) blocked the fatty acid-induced adhesion. However, unlike other systems, inhibition of ROCK did not block the activation of p38 mitogen-activated protein kinase (MAPK); instead, Rho activation depended on p38 MAPK activity and the presence of heat shock protein 27 (HSP27), which is phosphorylated downstream of p38 after arachidonic acid treatment. HSP27 associated with p115RhoGEF in fatty acid-treated cells, and this association was blocked when p38 was inhibited. Furthermore, siRNA knockdown of HSP27 blocked the fatty acid-stimulated Rho activity. Expression of dominant negative p115-RhoGEF or p115RhoGEF-specific siRNA inhibited both RhoA activation and adhesion on type IV collagen, whereas a constitutively active p115RhoGEF restored the arachidonic acid stimulation in cells in which the p38 MAPK had been inhibited. These data suggest that n-6 dietary fatty acids stimulate a set of interactions that regulates cell adhesion through RhoA and ROCK II via a p38 MAPK-dependent association of HSP27 and p115RhoGEF.The ability of tumor cells to metastasize to secondary sites is a hallmark of neoplastic disease. Unfortunately, this propensity to spread is the primary cause of morbidity and death in cancer patients (1). Metastasis is clearly a highly regulated, multistep process that occurs in a spatiotemporal manner (24). To escape the restrictive compartment boundaries characteristic of adult tissue, separate intravasation and extravasation steps requiring alterations in co-adhesion, adhesion, invasion, and migration must occur. Execution of these biological processes, involving multiple proteins and cellular organelles, require highly coordinated cell signaling mechanisms.The Rho family of small GTPases regulates many facets of cytoskeletal rearrangements that facilitate cell attachment and migration (57). Rho GTPases act as molecular switches by changing from an inactive GDP-bound conformation to an active GTP-bound conformation, thereby regulating a signaling pathway. These proteins are directly regulated by Rho guanine nucleotide exchange factors (GEFs),2 Rho GTPase activating proteins, and Rho GDP-dissociation inhibitors (812). RhoGEFs bind to the GTPase to catalyze the dissociation of GDP, allowing the binding of GTP and thereby promoting Rho activation (8). The RGS (regulators of G protein signaling) domain-containing RhoGEFs are a recently described family of GEFs. Currently, there are three members of this family, PDZ-RhoGEF, LARG, and p115RhoGEF (1315), in which the RGS domains function as a heterotrimeric GTPase-activating domain (13, 15, 16). The RGS family of RhoGEFs has been shown to regulate Rho during several processes including cytoskeletal rearrangements, cell adhesion, and cancer progression (1721).There is significant interplay between the activity of small GTPases and signaling derived from fatty acid metabolism (2228). Linoleic acid, which is metabolized to arachidonic acid, is an n-6 polyunsaturated fatty acid that is present at high levels in most western diets (29). In animal models, diets high in n-6 polyunsaturated fatty acids have been shown to enhance tumor progression and metastasis (30, 31). Additionally, arachidonic acid is stored in cell membranes and is made available by phospholipases under conditions of increased inflammatory response (32). Arachidonic acid is further metabolized by cyclooxygenases (COX), lipoxygenases (LOX), and cytochrome P450 monooxygenases to yield bioactive products that have myriad effects on cells, and altered metabolism of arachidonic acid by COX, LOX, and P450 has been implicated in cancer progression (31, 3336).We have studied mechanisms of cell adhesion using the MDA-MB-435 cells as a model of a highly metastatic human cancer cell line (37). These cells have been extensively studied for their ability to recapitulate the metastatic cascade in vivo and in vitro, although recent work indicates that the cells currently in use are most likely a human melanoma line (38). We initially observed that arachidonic acid (AA) enhanced adhesion of MDA-MB-435 cells to type IV collagen through specific integrin-mediated pathways (37). Exogenous AA led to the activation of mitogen-activated protein kinase (MAPK)-activated protein kinase 2 and the phosphorylation of heat shock protein 27 (HSP27) via a p38 MAPK-dependent process (39). Inhibition of p38 MAPK activation blocked cell adhesion as did function-blocking antibodies specific for subunits of the collagen receptor (40). More recently, we identified the key metabolite of AA (15-(S)- hydroxyeicosatetraenoic acid) and the upstream kinases (TAK1 and MKK6) that are responsible for activation of p38 MAPK in this system (41).In this study we investigated the role of Rho activation in the MDA-MB-435 cells after exposure to arachidonic acid. Several aspects of the regulation of Rho signaling in these cells provide insights into the cross-talk between important signaling pathways.  相似文献   

19.
20.
A central question in Wnt signaling is the regulation of β-catenin phosphorylation and degradation. Multiple kinases, including CKIα and GSK3, are involved in β-catenin phosphorylation. Protein phosphatases such as PP2A and PP1 have been implicated in the regulation of β-catenin. However, which phosphatase dephosphorylates β-catenin in vivo and how the specificity of β-catenin dephosphorylation is regulated are not clear. In this study, we show that PP2A regulates β-catenin phosphorylation and degradation in vivo. We demonstrate that PP2A is required for Wnt/β-catenin signaling in Drosophila. Moreover, we have identified PR55α as the regulatory subunit of PP2A that controls β-catenin phosphorylation and degradation. PR55α, but not the catalytic subunit, PP2Ac, directly interacts with β-catenin. RNA interference knockdown of PR55α elevates β-catenin phosphorylation and decreases Wnt signaling, whereas overexpressing PR55α enhances Wnt signaling. Taken together, our results suggest that PR55α specifically regulates PP2A-mediated β-catenin dephosphorylation and plays an essential role in Wnt signaling.Wnt/β-catenin signaling plays essential roles in development and tumorigenesis (13). Our previous work found that β-catenin is sequentially phosphorylated by CKIα4 and GSK3 (4), which creates a binding site for β-Trcp (5), leading to degradation via the ubiquitination/proteasome machinery (3). Mutations in β-catenin or APC genes that prevent β-catenin phosphorylation or ubiquitination/degradation lead ultimately to cancer (1, 2).In addition to the involvement of kinases, protein phosphatases, such as PP1, PP2A, and PP2C, are also implicated in Wnt/β-catenin regulation. PP2C and PP1 may regulate dephosphorylation of Axin and play positive roles in Wnt signaling (6, 7). PP2A is a multisubunit enzyme (810); it has been reported to play either positive or negative roles in Wnt signaling likely by targeting different components (1121). Toward the goal of understanding the mechanism of β-catenin phosphorylation, we carried out siRNA screening targeting several major phosphatases, in which we found that PP2A dephosphorylates β-catenin. This is consistent with a recent study where PP2A is shown to dephosphorylate β-catenin in a cell-free system (18).PP2A consists of a catalytic subunit (PP2Ac), a structure subunit (PR65/A), and variable regulatory B subunits (PR/B, PR/B′, PR/B″, or PR/B‴). The substrate specificity of PP2A is thought to be determined by its B subunit (9). By siRNA screening, we further identified that PR55α, a regulatory subunit of PP2A, specifically regulates β-catenin phosphorylation and degradation. Mechanistically, we found that PR55α directly interacts with β-catenin and regulates PP2A-mediated β-catenin dephosphorylation in Wnt signaling.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号